Next Article in Journal
Single Source Precursor for PAD-LaMnO3 Thin Films
Previous Article in Journal
A Multifaceted Kinetic Model for the Thermal Decomposition of Calcium Carbonate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Si Quantum Dots Assist Synthesized Microflower-Like Si/MoS2 Composites for Supercapacitors

1
School of Materials Science and Engineering, Chang’an University, Xi’an 710064, China
2
Engineering Research Central of Pavement Materials, Ministry of Education, Chang’an University, Xi’an 710064, China
*
Authors to whom correspondence should be addressed.
Crystals 2020, 10(9), 846; https://doi.org/10.3390/cryst10090846
Submission received: 15 August 2020 / Revised: 17 September 2020 / Accepted: 20 September 2020 / Published: 22 September 2020

Abstract

:
The microflower-like Si/MoS2 composites were fabricated using Si quantum dots (QDs) to assist a facile hydrothermal method. The electrochemical performance of Si/MoS2 composite in symmetric and asymmetric systems was studied. Electrochemical characterization revealed that the Si/MoS2 composite electrode in a three-electrode system has a high specific capacitance of 574.4 F·g−1 at 5 A·g−1. Furthermore, the Si/MoS2 composite electrode in a two-electrode system had the maximum energy density of 27.2 Wh·kg−1 when a power density of 749.1 W·kg−1 was achieved. Therefore, this investigation proves the Si/MoS2 composite microflower-like structure should be a promising candidate electrode material for supercapacitors.

Graphical Abstract

1. Introduction

The progress of science and technology has brought about rapid changes to human life, but has also brought about irreversible problems, such as global warming, the near exhaustion of non-renewable energy, and environmental pollution. Humans are eager to seek new, green energy to alleviate the energy crisis and environmental crisis, and even replace the traditional non-renewable energy [1,2,3]. Clean energy storage devices have been widely researched, due to their wider and wider application in numerous electronic devices. As a new type of clean energy storage device, supercapacitors have high power density, good cyclic stability, fast charge, and high rate of discharge [4,5,6,7,8].
Nowadays, two-dimensional (2D) materials have shown many unique advantages. Firstly, because electrons are confined to the 2D plane, especially for monolayer 2D materials, their electronic properties are enhanced. Therefore, 2D materials are ideal materials for basic research in condensed matter physics and electronic/optoelectronic devices. Secondly, because 2D materials can maintain the thickness of atoms while possessing the maximum plane size, they have a large specific surface area. This property has greatly attracted research into applications concerning surface area, such as catalysis and supercapacitors. Thirdly, ultra-thin 2D nanomaterials based on liquid phase processing can be prepared by simple methods into a single high-quality film, which is very necessary for practical applications, such as supercapacitors and solar cells.
2D transition metal disulfides (TMDs) have attracted widespread attention, due to their special layer structure. They have better performance of conductivity and larger surface area than oxide, the metal element and sulfur element through the weak van der Waals force of interaction between a single layer, forming a similar graphene layer structure [9]. This structure facilitates the insertion and extraction of various ions, and can be developed in the field of energy storage, including for lithium ion batteries and supercapacitors. Transition metal sulfides (MoS2 [10] CoS [11], NiS [12], CuS [13], etc.) are considered as potential materials for electrodes, because of their wide range of sources, low prices, and unique physicochemical properties. As a representative transition metal sulfide, MoS2 has received extensive attention in capacitor research [14]. Manuraj et al. [15] reported a specific capacitance of 972 F·g−1 at 1 A·g−1 for MoS2-RuO2 nanocomposite. Lien et al. [16] fabricated high voltage symmetric supercapacitors of graphene/MoS2, which works at a wide operating voltage of 2.3 V and achieves the maximum energy density of 31.2 Wh kg−1. Vattikuti et al. [17] reported that the hydrothermal reaction synthesized MoS2/MoO3 heterostructure electrodes have a high specific capacitance of 287.7 F·g−1 at 1 A·g−1 with good cycling stability. The MoS2 preparation method is simple and low cost. MoS2 has different morphologies, such as nanoball [18], nanoflake [19], nanoflower [20], nanofilament [21], nanoparticle [22], etc. It is found that MoS2 can not only show good capacitance in the electrochemical double layer, but can also generate an extra Faraday capacitance, due to the interaction between Li+, Na+, K+, and other ions inserted into the MoS2 layer. Nevertheless, the energy storage application of pure MoS2 as an electrode material is still very limited, with low conductivity and easy agglomeration, leading to a relatively low capacitance of MoS2. Therefore, the current research focus is to improve the electrochemical performance and increase the capacitance of composite materials
Zero-dimensional semiconductor nanomaterials can be sorted into three categories: IV group quantum dots, III-V group quantum dots [23], and II-VI group quantum dots. In the IV group, Si is one of the most important elements. Si, one of the most important elements, is the eighth most abundant element in the universe. Si nanomaterials are fully compatible with Si-based microelectronic devices, and nanosilicates have many unusual optical and electrical properties [24]. As a result of high research value, both in the basic theory and practical applications, Si nanomaterials have attracted considerable attention. The quantum confinement effect occurs when the particle size of a Si nanocrystal is reduced to the radius of the exciton Bohr. Si nanocrystals are then called Si QDs. In Si QDs, the motion of electrons or holes in three-dimensional space is constrained by the fact that the constrained carrier can only be located at the separated bound level, and the motion is fully quantized, thus weakening the constraint of momentum conservation. Therefore, the electronic structure of Si semiconductor nanocrystals is different from that of bulk materials of the same composition. It is mainly manifested in the enhancement of electron–hole exchange, the increase of exciton binding energy, and the increase of optical transition oscillation intensity with the decrease of semiconductor gap width. Among the quantum constraint effects, the most reported is the quantum size effect, that is, the band gap increases as the quantum dot size decreases. Si QDs have been developed in recent years with many advantages. The low preparation cost lays a foundation for the rapid application of Si QDs in production. Its low toxicity provides a security guarantee for the experimental research of Si QDs [25]. Its chemical properties provide a basis for the study of its electrochemical properties.
In this article, we propose a scheme for the synthesis of MoS2 using Si QDs to assist hydrothermal synthesis. Using the Si QDs as the center of nucleation promotes the nucleation process of MoS2, and accelerates the reaction speed. Si/MoS2 composite presents a flower-like structure, with the petals crisscrossing each other. This structure makes the active substance fully contact the electrolyte, which results in the redox reaction. Moreover, to a certain extent, the structure reduces the changes in the microstructure of the material caused by the charging and discharging process. By comparing the electrochemical performance of the two kinds of electrodes (MoS2 and Si/MoS2 composite), the Si/MoS2 composite electrode demonstrates excellent performance, including high specific capacitance and good cyclic stability. This work can arouse wide interest in the preparation of high-performance mixed-metal sulfide electrode materials by a simple method.

2. Materials and Methods

A total of 0.1 g of sodium ascorbate was dissolved in 10 mL deionized (DI) water while stirring for 30 min, then, a solution could be obtained. Following this, 3.75 mL of sodium ascorbate solution, 1 mL of N-aminoethyl-γ-aminopropyltrimethoxysilane, and 12 mL DI water were mixed with magnetic stirring for 20 min to obtain a Si quantum dot solution.
Firstly, 0.7062 g (NH4)6Mo7O24·6H2O and 1.3049 g H2NCSNH2 were diluted with 20 mL DI water. The prepared solution was mixed with 1 mL Si QD solution, and subjected to continuous magnetic stirring for 1 h. Subsequently, the reaction mixture was kept in a Teflon lined stainless steel autoclave at 180 °C for 18 h. After that, the final mixture product was filtered, washed with DI water, and dried at 60 °C in a vacuum oven to obtain the Si/MoS2 composite. For better comparison, the MoS2 was prepared following the same procedures, but without the Si QDs.
An electrochemical station (CHI 660E, Chenhua, Shanghai, China) was taken to test electrochemical performance in 6 M KOH. The sample was used as the working electrode, with a platinum foil (2 cm × 2 cm) and a saturated calomel electrode as the counter and the reference electrode, respectively. The specific capacitance of the single electrode was calculated using the equation:
C = i Δ t / ( S Δ V )
C = i Δ t / ( m Δ V )
where C is the specific capacitance, i is the discharge current, Δt is the discharge time, S and m are assigned as the area and mass of the electrode material, and ΔV represents the potential window.

3. Results

Fourier transform infrared (FT-IR) spectroscopy (TENSOR II, Bruker, Karlsruhe, Germany) was used to investigate the bonding interaction and functional groups of samples [26]. As shown in Figure 1a, the peak appearing at 590 cm−1 is assigned to the Mo–S vibration [27], the peak at 1010 cm−1 corresponds to the S–O stretching [28], and the peak at 1413 cm−1 is attributed to C–O stretching vibration [29]. For MoS2, the peaks observed at 1132 and 1651 cm−1 are attributed to the C–H bending vibration [27] and the existence of C=O stretching vibration [30]. For Si/MoS2, the absorption peak at 1106 cm-1 is ascribed to the characteristic stretching of the Si–O–Si band [31]. Figure 1b gives the XRD pattern for Si/MoS2 composite. The main diffraction peaks of MoS2 at 2θ = 33.3°, 39.2° and 48.5° correspond to (101), (103), and (105) reflections, respectively (JCPDS card no. 37-1492). The peak of Si at 2θ = 58.6° can be attributed to the (101) plane (JCPDS card no. 47-1186).
X-ray photoelectron spectroscopy (XPS, ESCA-LAB Mk II, VG Scientific Co., London, England) was used to analyze the elemental valence state and surface chemical composition. Figure 2 shows the XPS high-resolution spectra of Mo 3d and S 2p. The peaks at 230.3 eV, 233.5 eV, and 236.3 eV, observed in the pristine spectrum of Mo in Figure 2a, are identified as Mo4+ 3d5/2, Mo4+ 3d3/2, and Mo6+ 3d3/2, respectively [32]. The presence of Mo6+ indicates that there may be a small amount of MoO3 in the sample [33]. Moreover, there is another peak at 227.1 eV, which represents S2- 2s [34]. For S 2p in Figure 2b, the peaks at 163.7 eV and 164.2 eV, observed in the pristine spectrum of S, correspond to S 2p5/2 and S 2p3/2.
In order to study the effect of Si QDs on the microstructure of the whole composite, field emission scanning electron microscope (FESEM, S-4800, Hitachi, Tokyo, Japan) analysis was used for MoS2 and Si/MoS2 composite. Figure 3 shows the SEM images of MoS2 and Si/MoS2 composite. In Figure 3a–c, it can be seen that without Si QDs, MoS2 is composed of irregular nanosheets with a diameter of 200–300 nm. Most of the nanosheets are tiled, and there are some curled-off phenomena. On the other hand, it still conforms to the rule that MoS2 nanosheets tend to grow in flat under the control of MoS2. Figure 3d,f is the SEM diagram of Si/MoS2 composite material. After adding a small amount of Si QDs, the Si/MoS2 composite presents a microflower spherical structure composed of staggered nanosheets, which indicates that the MoS2 nanosheets grow with Si QDs as nucleation centers, and finally grow into three-dimensional flower spheres. The energy dispersive X-ray (EDX) elemental maps in Figure 3g–i reveal that there are Si, Mo, and S elements in the Si/MoS2 composite. It is evident that the Si QDs are uniformly distributed in Si/MoS2 composite.
A possible growth mechanism is depicted in the schematic diagram in Figure 4. Without Si QDs, MoS2 tends to grow tiled in the direction with the minimum binding potential energy. With the prolongation of reaction time, the MoS2 nanosheets grew more and more, and were stacked. A small part of MoS2 nanosheets can overcome the crimping of partial binding potential energy or the semi-upright growth. It may be caused by the dynamic force generated by the easy agglomeration of MoS2 itself, which can overcome the crimp growth of certain binding potential energy, and finally form the structure in Figure 3c; when, with Si QDs, due to the small size of Si QDs, it can be used as the nucleation center in the reaction process. The generated MoS2 nanosheet tends to pile up around Si QDs, overcoming the potential energy for rapid three-dimensional growth, and finally forming a micron flower ball. As a result, the structure in Figure 3f is formed with the continuation of the reaction time.
Figure 5a is the transmission electron microscope (TEM, Tecnai F30G2, FEI, Hillsboro, OR, USA) image of Si/MoS2. It confirms the microflower structure further, in which many nanosheets are stacked on each other. The result is consistent with the SEM image in Figure 3f. Figure 5b is a high-resolution-TEM image of Si/MoS2, in which it is clearly observable the ordered lattice fringes with the corresponding interplanar distance of 0.61 nm, which relates to the (002) plane of the MoS2 [35]. Figure 5c is the high-resolution image after the Fourier transform of the box area in Figure 5b. The crystal lattice of the sample can be observed to be 0.22 nm, which corresponds to the (100) plane of the MoS2.
In order to compare the electrochemical properties of the MoS2 and Si/MoS2 composite, CV curves with a scan rate of 100 mV·s−1 were given in Figure 6a,b. In Figure 6a, the integral areas of the CV curve of the Si/MoS2 composite are bigger than those of MoS2, which suggests that the specific capacitance of Si/MoS2 composite is bigger than the MoS2. The GCD curves of the MoS2 and Si/MoS2 composite at the same current density of 50 A·g−1 are shown in Figure 6b. The plateaus in the GCD curves indicate that the two samples performed pseudocapacitor behaviors [36], which is consistent with the CV results. In Figure 6b, Si/MoS2 composite possesses a longer discharge time than MoS2, which indicates that it has a large specific capacitance. The specific capacitance of Si/MoS2 composite is higher than that of MoS2 for the following three reasons: (1) after the addition of Si QDs, Si QDs as the nucleation center, play the role of supporting MoS2 nanosheets, which greatly increases the ion transport channel; (2) the porous and open microflower structure of Si/MoS2 composite, which can shorten the diffusion path of electrons and ions and enhance electrochemical dynamics [37]; and (3) the open space of the Si/MoS2 composite microflower can act as the robust reservoir for electrolyte ions, supplying sufficient redox reactions for energy storage [38]. The CV curves for MoS2 and Si/MoS2 composite at different scan rates of 5–100 mV/s are shown in Figure 6c,d. With the increase of the scanning rate, the peak of current gradually increased, and the curve shape remained basically unchanged in the CV curves of MoS2 and Si/MoS2 composite, indicating that the two electrode materials had the potential to provide high power performance. The GCD curves of the MoS2 and Si/MoS2 composite at different current densities are shown in Figure 6e,f. The specific capacitances of the MoS2 and Si/MoS2 composite can be calculated on the basis of the GCD curves, and the results are presented in Figure 6g. The specific capacitance of Si/MoS2 composite at any of the same current densities is higher than that of MoS2. With the increase of current density, the discharge time is shortened continuously, so the specific capacity is also decreased continuously. When the current density of Si/MoS2 composite is at the 5 A·g−1, 10 A·g−1, 20 A·g−1, 30 A·g−1, 40 A·g−1, and 50 A·g−1, the corresponding specific capacitance of Si/MoS2 composite electrode is 574.4 F·g−1, 500.0 F·g−1, 404.4 F·g−1, 284.4 F·g−1, and 233.3 F·g−1, respectively, which is consistent with the analysis results of CV curves in Figure 6c,d. Figure 6i shows cyclic stability of capacitive performance of Si/MoS2 electrode tested for 1000 cycles at 1 A·g−1. The Si/MoS2 composite electrode exhibits specific capacitance retention of 84.5% for 1000 cycles.
Nyquist plots of the MoS2 and Si/MoS2 composite electrodes are shown in Figure 6h. Generally, a semicircle in the high frequency region and a straight line with slope in the low frequency region are shown. In the low frequency region, it consists of a straight line, while in the high frequency region, it consists of a semicircle. The diameter of the semicircle is related to the charge transfer resistance (Rct). The radius of the Rct from the high frequency arc is on the real axis. Additionally, the smaller the radius, the smaller the Rct. A smaller Rct for Si/MoS2 and a vertical line along the virtual axis show lower resistance and better capacitive behavior [39]. The Nyquist curve of the X-ray intercepts the equivalent series resistance of said electrode (RS), which is 0.506 Ω.
In order to evaluate the potential practical application of the Si/MoS2 composite electrode, an asymmetric capacitor (Si/MoS2//AC) was constructed, with Si/MoS2 composite active material as positive electrode and AC as negative electrode, to investigate its electrochemical performance. An electrochemical workstation was used to test the asymmetric supercapacitor electrochemical properties. As shown in Figure 7a, CV curves of the Si/MoS2 composite and AC separately tested in the three-electrode test system indicated that the optimal potential window of Si/MoS2//AC is 1.5 V, and a good matching for asymmetric supercapacitors. Figure 7b shows CV curves of Si/MoS2//AC at various scan rates, which show that the pseudocapacitance and double-layer capacitance are common influences on the supercapacitor. The shapes of CV curves were not deformed when the sweep speeds increased, indicating good rate performance. The C (specific capacitance (86.9 F·g−1)) in the asymmetric capacitor was calculated based on the Figure 7c discharge curves when current density was at 1 A·g−1. The Ragone plot is displayed in Figure 7d. The maximum energy density of 27.2 Wh·kg−1 is obtained at a power density of 749.1 Wh·kg−1, which is more than other reported systems, such as MoS2-150 mg GF//AEG (16 Wh·kg−1/758.2 W·kg−1) [6], BCN/MoS2-31 composite (8.92 Wh·kg−1/254.8 W·kg−1) [40], PANI-G-MoS2 (2.65 Wh·kg−1/119.2 W·kg−1) [41], and MoS2/RCF electrode (22.5 Wh·kg−1/703 W·kg−1) [9].

4. Conclusions

In this work, microflower-like structure Si/MoS2 composite was prepared by hydrothermal method. The addition of Si QDs can improve the electrochemical performance of MoS2. The capacitance of the Si/MoS2 composite is 574.4 F·g−1 at 5 A·g−1. When the energy density is 27.2 Wh·kg−1, the power density of the Si/MoS2//AC reaches 749.1 W·kg−1. The capacitance retention is 84.5%, even after 1000 cycles. Therefore, the Si/MoS2 composite microflower will be a great potential electrode material for supercapacitors.

Author Contributions

J.Z. and R.Z. conceived and designed the experiments; R.Z. and K.C. fabricated the sample; Y.Y. and Y.W. characterized the sample; P.Y. oversaw project administration. All authors have read and agreed to the published version of the manuscript.

Funding

The work was funded by National Natural Science Foundation of China (Grant No. 21607013, 5102026), the Natural Science Basic Research Plan in Shaanxi Province of China (2020JM-221), the Fund Project of Shaanxi Key Laboratory of Land Consolidation (2018-JC01), the Fundamental Research Funds for the Central Universities, CHD (300102310203), and Colation and Entrepreneurship Project (201810710127, 201910710466).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fu, H.; Zhang, X.D.; Fu, J.Z.; Shen, G.Z.; Ding, Y.L.; Chen, Z.J.; Du, H. Single layers of MoS2/Graphene nanosheets embedded in activated carbon nanofibers for high-performance supercapacitor. J. Alloy Compd. 2020, 829, 154557. [Google Scholar] [CrossRef]
  2. Wang, C.; Xu, J.; Yuen, M.; Zhang, J.; Li, Y.; Chen, X.; Zhang, W. Hierarchical composite electrodes of nickel oxide nanoflake 3D graphene for high-performance pseudocapacitors. Adv. Funct. Mater. 2014, 24, 6372–6380. [Google Scholar] [CrossRef]
  3. Reddy, B.P.; Sekhar, M.C.; Vattikuti, S.V.P.; Suh, Y.; Park, S.H. Solution-based spin-coated tin sulfide thin films for photovoltaic and supercapacitor applications. Mater. Res. Bull. 2018, 103, 13–15. [Google Scholar] [CrossRef]
  4. Vikraman, D.; Karuppasamy, K.; Hussain, S.; Kathalingam, A.; Sanmugam, A.; Jung, J.; Kim, H.S. One-pot facile methodology to synthesize MoS2-graphene hybrid nanocomposites for supercapacitors with improved electrochemical capacitance. Compos. Part B Eng. 2019, 161, 555–563. [Google Scholar] [CrossRef]
  5. Vattikuti, S.V.P.; Reddy, B.P.; Byon, C.; Shim, J. Carbon/CuO nanosphere-anchored g-C3N4 nanosheets as ternary electrode material for supercapacitors. J. Solid State Chem. 2018, 262, 106. [Google Scholar] [CrossRef]
  6. Masikhwa, T.M.; Madito, M.J.; Bello, A.; Dangbegnon, J.K.; Manyala, N. High performance asymmetric supercapacitor based on molybdenum disulphide/graphene foam and activated carbon from expanded graphite. J. Colloid Interf. Sci. 2017, 488, 155–165. [Google Scholar] [CrossRef]
  7. Yadav, S.K.; Kumar, R.; Sundramoorthy, A.K.; Singh, R.K.; Koo, C.M. Simultaneous reduction and covalent grafting of polythiophene on graphene oxide sheets for excellent capacitance retention. RSC Adv. 2016, 6, 52945–52949. [Google Scholar] [CrossRef] [Green Version]
  8. Vattikuti, S.V.P.; Devarayapalli, K.C.; Nagajyothi, P.C.; Shim, J. Microwave synthesized dry leaf-like mesoporous MoSe2 nanostructure as an efficient catalyst for enhanced hydrogen evolution and supercapacitor applications. Microchem. J. 2020, 153, 104446. [Google Scholar] [CrossRef]
  9. Zhao, C.H.; Zhou, Y.N.; Ge, Z.X.; Zhao, C.G.; Qian, X.Z. Facile construction of MoS2/RCF electrode for high-performance supercapacitor. Carbon 2018, 127, 699–706. [Google Scholar] [CrossRef]
  10. Xu, C.Y.; Jiang, L.; Li, X.; Li, C.; Shao, C.X.; Zuo, P.; Liang, M.S.; Qu, L.T.; Cui, T.H. Miniaturized high-performance metallic 1T-Phase MoS2 micro-supercapacitors fabricated by temporally shaped femtosecond pulses. Nano Energy 2020, 67, 104260. [Google Scholar] [CrossRef]
  11. Lia, J.F.; Chen, D.D.; Wu, Q.S. Facile synthesis of CoS porous nanoflake for high performance supercapacitor electrode materials. J. Energy Storage 2019, 23, 511–514. [Google Scholar] [CrossRef]
  12. Harish, S.; Naveen, A.N.; Abinaya, R.; Archana, J.; Ramesh, R.; Navaneethan, M.; Shimomura, M.; Hayakawa, Y. Enhanced performance on capacity retention of hierarchical NiS hexagonal nanoplate for highly stable asymmetric supercapacitor. Electrochim. Acta 2018, 283, 1053–1062. [Google Scholar] [CrossRef]
  13. Himasree, P.; Durga, I.K.; Krishna, T.N.V.; Rao, S.S.; Gopi, C.V.V.M.; Revathi, S.; Prabakar, K.; Kim, H.J. One-step hydrothermal synthesis of CuS@MnS on Ni foam for high performance supercapacitor electrode material. Electrochim. Acta 2019, 305, 467–473. [Google Scholar] [CrossRef]
  14. Kumar, R.; Sahoo, S.; Joanni, E.; Singh, R.K.; Yadav, R.M.; Verma, R.K.; Singh, D.P.; Tan, W.K.; del Pino, A.P.; Moshkalev, S.A.; et al. A review on synthesis of graphene, h-BN and MoS2 for energy storage applications: Recent progress and perspectives. Nano Res. 2019, 12, 2655–2694. [Google Scholar] [CrossRef]
  15. Manuraj, M.; Chacko, J.; Unni, K.N.N.; Rakhi, R.B. Heterostructured MoS2-RuO2 nanocomposite: A promising electrode material for supercapacitors. J. Alloy Compd. 2020, 836, 155420. [Google Scholar] [CrossRef]
  16. Lien, C.W.; Vedhanarayanan, B.; Chen, J.H.; Lin, J.Y.; Tsai, H.H.; Shao, L.D.; Lin, T.W. Optimization of acetonitrile/water content in hybrid deep eutectic solvent for graphene/MoS2 hydrogel-based supercapacitors. Chem. Eng. J. 2021, 405, 126706. [Google Scholar] [CrossRef]
  17. Vattikuti, S.V.P.; Nagajyothi, P.C.; Reddy, P.A.K.; Kumar, M.K.; Shim, J.; Byon, C. Tiny MoO3 nanocrystals self-assembled on folded molybdenum disulfide nanosheets via a hydrothermal method for supercapacitor. Mater. Res. Lett. 2018, 6, 432–439. [Google Scholar] [CrossRef] [Green Version]
  18. Feng, D.; Pan, X.X.; Xia, Q.Y.; Qin, J.H.; Zhang, Y.; Chen, X.M. Metallic MoS2 nanosphere electrode for aqueous symmetric supercapacitors with high energy and power densities. J. Mater. Sci. 2020, 55, 713–723. [Google Scholar] [CrossRef]
  19. Voiry, D.; Salehi, M.; Silva, R.; Fujita, T.; Chen, M.W.; Asefa, T.; Shenoy, V.B.; Eda, G.; Chhowalla, M. Conducting MoS2 Nanosheets as Catalysts for Hydrogen Evolution Reaction. Nano Lett. 2013, 13, 6222−6227. [Google Scholar] [CrossRef]
  20. Mishra, S.; Maurya, P.K.; Mishra, A.K. 2H–MoS2 nanoflowers based high energy density solid state supercapacitor. Mater. Chem. Phys. 2020, 255, 123551. [Google Scholar] [CrossRef]
  21. Wu, D.; Lou, Z.; Wang, Y.; Yao, Z.; Xu, T.; Shi, Z.; Xu, J.; Tian, Y.; Li, X.; Tsang, Y.H. Photovoltaic high-performance broadband photodetector based on MoS2/Si nanowire array heterojunction. Sol. Energ. Mat. Sol. C. 2018, 182, 272–280. [Google Scholar] [CrossRef]
  22. Feng, Z.; Yang, P.P.; Wen, G.S.; Li, H.B.; Liu, Y.; Zhao, X.C. One-step synthesis of MoS2 nanoparticles with different morphologies for electromagnetic wave absorption. Appl. Surf. Sci. 2020, 502, 144129. [Google Scholar] [CrossRef]
  23. Kumar, R.; Singh, R.K.; Yadav, S.K.; Savu, R.; Moshkalev, S.A. Mechanical pressure induced chemical cutting of boron nitride sheets into boron nitride quantum dots and optical properties. J. Alloy. Compd. 2016, 683, 38–47. [Google Scholar] [CrossRef]
  24. Singh, R.K.; Kumar, R.; Singh, D.P.; Savu, R.; Moshkalev, S.A. Progress in microwave-assisted synthesis of quantum dots (graphene/carbon/semiconducting) for bioapplications: A review. Mater. Today Chem. 2019, 12, 282–332. [Google Scholar] [CrossRef]
  25. Erogbogbo, F.; Yong, K.T.; Roy, I.; Hu, R.; Law, W.C.; Zhao, W.W.; Ding, H.; Wu, F.; Kumar, R.; Swihart, M.T.; et al. In vivo targeted cancer imaging, sentinel lymph node mapping and multi-channel imaging with biocompatible silicon nanocrystals. ACS Nano 2011, 5, 413–423. [Google Scholar] [CrossRef]
  26. Arefi-Oskoui, S.; Khataee, A.; Safarpour, M.; Vatanpour, V. Modification of polyethersulfone ultrafiltration membrane using ultrasonicassisted functionalized MoS2 for treatment of oil refinery wastewater. Sep. Purif. Technol. 2020, 238, 116495. [Google Scholar] [CrossRef]
  27. Zhang, X.L.; Ma, L.; Gan, M.Y.; Fu, G.; Jin, M.; Zhai, Y.F. Controllable constructing of hollow MoS2/PANI core/shell microsphere for energy storage. Appl. Surf. Sci. 2018, 460, 48–57. [Google Scholar] [CrossRef]
  28. Raza, A.; Ikram, M.; Aqeel, M.; Imran, M.; Ul-Hamid, A.; Riaz, K.N.; Ali, S. Enhanced industrial dye degradation using Co doped in chemically exfoliated MoS2 nanosheets. Appl. Nanosci. 2020, 10, 1535–1539. [Google Scholar] [CrossRef]
  29. Zhong, Y.; Sun, X.; Wang, S.; Peng, F.; Bao, F.; Su, Y.; Li, Y.; Lee, S.T.; He, Y. Facile, Large-Quantity Synthesis of Stable, Tunable-Color Silicon Nanoparticles and Their Application for Long-Term Cellular Imaging. ACS Nano 2015, 9, 5958–5959. [Google Scholar] [CrossRef]
  30. Liu, Z.P.; Hou, J.Z.; Wang, X.F.; Hou, C.J.; Ji, Z.; He, Q.; Huo, D.Q. A novel fluorescence probe for rapid and sensitive detection of tetracyclines residues based on silicon quantum dots. Spectrochim. Acta A 2020, 240, 118463. [Google Scholar] [CrossRef]
  31. Xu, X.X.; Jiao, S.L.; Liu, Z.Q.; Liu, L. Synergistic lubrication of a porous MoS2-POSS nanohybrid. RSC Adv. 2020, 10, 20579. [Google Scholar] [CrossRef]
  32. Yang, Z.M.; Zhang, D.Z.; Chen, H.N. MOF-derived indium oxide hollow microtubes/MoS2 nanoparticles for NO2 gas sensing. Sensor. Actuat. B Chem. 2019, 300, 127037. [Google Scholar] [CrossRef]
  33. Zhao, P.; Tang, Y.; Mao, J.; Chen, Y.; Song, H.; Wang, J.; Song, Y.; Liang, Y.; Zhang, X. One-dimensional MoS2-decorated TiO2 nanotube gas sensors for efficient alcohol sensing. J. Alloy. Compd. 2016, 674, 252–258. [Google Scholar] [CrossRef]
  34. Zhao, S.; Li, Z.; Wang, G.; Liao, J.; Lv, S.; Zhu, Z. Highly enhanced response of MoS2/porous silicon nanowire heterojunctions to NO2 at room temperature. RSC Adv. 2018, 8, 11070–11077. [Google Scholar] [CrossRef] [Green Version]
  35. Luo, J.H.; Zhang, K.; Cheng, M.L.; Gu, M.M.; Sun, X.K. MoS2 spheres decorated on hollow porous ZnO microspheres with strong wideb and microwave absorption. Chem. Eng. J. 2020, 380, 122625. [Google Scholar] [CrossRef]
  36. Arasi, S.E.; Ranjithkumar, R.; Devendran, P.; Krishnakumar, M.; Arivarasan, A. Electrochemical evaluation of binary Ni2V2O7 nanorods as pseudocapacitor electrode material. Ceram. Int. 2020, 46, 22709–22717. [Google Scholar] [CrossRef]
  37. Guan, B.; Li, Y.; Yin, B.Y.; Liu, K.F.; Wang, D.W.; Zhang, H.H.; Cheng, C.J. Synthesis of hierarchical NiS microflower for high performance asymmetric supercapacitor. Chem. Eng. J. 2017, 308, 1165–1173. [Google Scholar] [CrossRef]
  38. Huo, H.H.; Zhao, Y.Q.; Xu, C.L. 3D Ni3S2 nanosheet arrays supported on Ni foam for high-performance supercapacitor and nonenzymatic glucose detection. J. Mater. Chem. A 2014, 2, 15111–15117. [Google Scholar] [CrossRef]
  39. Peng, W.J.; Wang, W.; Han, G.H.; Huang, Y.K.; Zhang, Y.S. Fabrication of 3D flower-like MoS2/graphene composite as highperformance electrode for capacitive deionization. Desalination 2020, 473, 114191. [Google Scholar] [CrossRef]
  40. Thakur, A.K.; Majumder, M.; Choudhary, R.B.; Singh, S.B. MoS2 flakes integrated with boron and nitrogen-doped carbon: Striking gravimetric and volumetric capacitive performance for supercapacitor applications. J. Power Sources 2018, 402, 163–173. [Google Scholar] [CrossRef]
  41. Palsaniya, S.; Nemade, H.B.; Dasmahapatra, A.K. Synthesis of polyaniline/graphene/MoS2 nanocomposite for high performance supercapacitor electrode. Polymer 2018, 150, 150–158. [Google Scholar] [CrossRef]
Figure 1. FT-IR spectrum of MoS2 and Si/MoS2 composite electrodes (a), and XRD pattern of Si/MoS2 composite (b).
Figure 1. FT-IR spectrum of MoS2 and Si/MoS2 composite electrodes (a), and XRD pattern of Si/MoS2 composite (b).
Crystals 10 00846 g001
Figure 2. The XPS spectra of (a) the Mo 3d; and (b) the S 2p.
Figure 2. The XPS spectra of (a) the Mo 3d; and (b) the S 2p.
Crystals 10 00846 g002
Figure 3. SEM images of MoS2 (ac), S and Si/MoS2 composite (df), and EDX elemental mapping image of Si (g), Mo (h), and S (i) in Si/MoS2 composite.
Figure 3. SEM images of MoS2 (ac), S and Si/MoS2 composite (df), and EDX elemental mapping image of Si (g), Mo (h), and S (i) in Si/MoS2 composite.
Crystals 10 00846 g003
Figure 4. Schematic illustration of the formation for MoS2 and Si/MoS2 composite.
Figure 4. Schematic illustration of the formation for MoS2 and Si/MoS2 composite.
Crystals 10 00846 g004
Figure 5. TEM image of Si/MoS2 composite (a); high-resolution TEM images (b) and (c).
Figure 5. TEM image of Si/MoS2 composite (a); high-resolution TEM images (b) and (c).
Crystals 10 00846 g005
Figure 6. (a,b) Comparison of CV and GCD curves of two electrodes; (c,e) CV and GCD curves of MoS2 and (d,f) Si/MoS2 composite; (g) rate capacitance calculated by GCD curves; (h) EIS curves of two electrodes (the inset shows the equivalent circuit model used for fitting the EIS plot); (i) cyclic stability test at 1 A·g−1 for Si/MoS2 composite electrode.
Figure 6. (a,b) Comparison of CV and GCD curves of two electrodes; (c,e) CV and GCD curves of MoS2 and (d,f) Si/MoS2 composite; (g) rate capacitance calculated by GCD curves; (h) EIS curves of two electrodes (the inset shows the equivalent circuit model used for fitting the EIS plot); (i) cyclic stability test at 1 A·g−1 for Si/MoS2 composite electrode.
Crystals 10 00846 g006aCrystals 10 00846 g006b
Figure 7. (a) CV curve of AC and Si/MoS2 composite, separately tested in three-electrode system at 100 mV·s−1; (b) CV curve of asymmetric supercapacitor (Si/MoS2//AC); (c) GCD curve of Si/MoS2//AC; (d) Ragone plots of Si/MoS2//AC.
Figure 7. (a) CV curve of AC and Si/MoS2 composite, separately tested in three-electrode system at 100 mV·s−1; (b) CV curve of asymmetric supercapacitor (Si/MoS2//AC); (c) GCD curve of Si/MoS2//AC; (d) Ragone plots of Si/MoS2//AC.
Crystals 10 00846 g007

Share and Cite

MDPI and ACS Style

Zheng, J.; Cheng, K.; Zhang, R.; Yang, Y.; Wu, Y.; Yu, P. Si Quantum Dots Assist Synthesized Microflower-Like Si/MoS2 Composites for Supercapacitors. Crystals 2020, 10, 846. https://doi.org/10.3390/cryst10090846

AMA Style

Zheng J, Cheng K, Zhang R, Yang Y, Wu Y, Yu P. Si Quantum Dots Assist Synthesized Microflower-Like Si/MoS2 Composites for Supercapacitors. Crystals. 2020; 10(9):846. https://doi.org/10.3390/cryst10090846

Chicago/Turabian Style

Zheng, Jiahong, Kangkang Cheng, Runmei Zhang, Yamei Yang, Yuntao Wu, and Pengfei Yu. 2020. "Si Quantum Dots Assist Synthesized Microflower-Like Si/MoS2 Composites for Supercapacitors" Crystals 10, no. 9: 846. https://doi.org/10.3390/cryst10090846

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop